Mono-, Di-, and Trifluoroalkyl Substituent Effects on the

Aug 24, 2015 - The reactivities and torquoselectivities of electrocyclic ring opening reactions of fluoromethyl-substituted cyclobutenes and oxetenes ...
0 downloads 0 Views 1MB Size
Article pubs.acs.org/joc

Mono‑, Di‑, and Trifluoroalkyl Substituent Effects on the Torquoselectivities of Cyclobutene and Oxetene Electrocyclic Ring Openings Kazuya Honda,†,§ Steven A. Lopez,‡,§ K. N. Houk,*,‡ and Koichi Mikami*,† †

Department of Applied Chemistry, Tokyo Institute of Technology, Tokyo 152-8552, Japan Department of Chemistry & Biochemistry, University of California, Los Angeles, California 90095-1569, United States



Downloaded by UNIV OF NEBRASKA-LINCOLN on September 2, 2015 | http://pubs.acs.org Publication Date (Web): August 28, 2015 | doi: 10.1021/acs.joc.5b01361

S Supporting Information *

ABSTRACT: The reactivities and torquoselectivities of electrocyclic ring opening reactions of fluoromethyl-substituted cyclobutenes and oxetenes were studied with M06-2X density functional theory. The torquoselectivities of a series of mono-, di-, and trifluoromethylcyclobutenes and oxetenes result from the interplay of favorable orbital interactions and closed-shell repulsions. When the substituent rotates inward, there can be a favorable interaction between the breaking σCO bond and the σCF* orbital (σCO → σCF*) of the fluoromethyl group in fluoromethyloxetenes. The preference for rotation of a fluoromethyl group is decreased in trifluoromethyloxetenes because closed-shell repulsions between the breaking σCO bond and trifluoromethyl substituent orbitals compete with the σCO → σCF* interaction.



stereoelectronic factors may be definitive.”1 They showed the example of a ring opening that occurs upon solvolysis of cyclopropyl tosylates and related species. The leaving groups dictate which disrotatory process occurs.1 In 1984, our group explored the selectivity of the conrotatory ring openings of substituted cyclobutenes.2c Torquoselectivity is “the preference for inward or outward rotation of substituents in conrotatory or disrotatory electrocyclic reactions”, the selectivity as “preferentially along one of two [allowed] paths.” In hundreds of papers since then, our groups and others have established the robustness of the torquoselectivity rules. In short, while large groups prefer to rotate outward, as Woodward and Hoffmann anticipated, electronic factors are even more important: π-donors rotate outward, and πacceptors rotate inward. Cyclobutenes undergo thermal conrotatory 4π electrocyclic ring opening reactions to afford 1,3-butadienes.1 Substituents

INTRODUCTION One of us (K.N. Houk) joined R.B. Woodward’s research group as a first-year graduate student at Harvard in early January of 1965, just when Woodward and Hoffmann published the JACS communication, “Stereochemistry of Electrocyclic Reactions”. In that communication, Woodward and Hoffmann defined conrotatory and disrotatory stereoselectivities and gave theoretical insights into the rules that show which stereochemistry is favored, based upon the number of electrons involved. There have been many theoretical elaborations of these rules by correlation diagrams, concepts of aromaticity, and frontier molecular orbitals. This aspect of Woodward− Hoffmann theory remains the most robust aspect, and experimentalists have shown as much as 15 kcal mol−1 differences between allowed and forbidden activation barriers. In the last paragraph of the 1965 communication, Woodward and Hoffmann gave one example of what our group later named “torquoselectivity”, the preference for one of the allowed modes of electrocyclic ring opening. Woodward and Hoffmann noted “Ordinarily, simple steric factors will be expected to direct the changes preferentially along one of the two paths, but in some cases, very interesting special © XXXX American Chemical Society

Special Issue: 50 Years and Counting: The Woodward-Hoffmann Rules in the 21st Century Received: June 24, 2015

A

DOI: 10.1021/acs.joc.5b01361 J. Org. Chem. XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF NEBRASKA-LINCOLN on September 2, 2015 | http://pubs.acs.org Publication Date (Web): August 28, 2015 | doi: 10.1021/acs.joc.5b01361

The Journal of Organic Chemistry may rotate “inward’ or “outward” during the course of reaction; a preference for one of these diastereomeric transition states is called torquoselectivity.2 This selectivity has been shown to arise from interactions between the substitutent orbitals and those of the breaking bond. Donors rotate outward to avoid repulsive filled−filled interactions with the HOMO of cyclobutene upon inward rotation. Acceptors have low-lying vacant orbitals and the best acceptors (CHO, − SiR3, − GeR3) preferentially rotate inward. The reactivities and torquoselectivities of four-membered heterocycles such as 2-azetines,3 oxetenes, and thietes4 have received renewed attention. These cyclobutenes and heterocyclic derivatives can be designed to undergo ring opening reactions to afford functionalized dienes and heterodienes stereoselectively. The trifluoromethyl group is of great importance in synthetic organic chemistry5 and medicinal6 chemistry because of its unique steric and electronic properties.7 While major strides have been made in late-stage functionalization of heteroaromatic substrates with the CF3 group, stereoselective functionalization of tri- and tetrasubstituted alkenes is quite limited.8 Recent work by Mikami et al. fills this void in the literature by utilizing a torquoselective electrocyclic ring opening reaction of oxetenes9 (Scheme 1) inspired by earlier reports by Houk et al. on the electrocyclic ring opening of 3-trifluoromethylcyclobutene.10

rotates inward. The favorable (σCO → σCF*) interaction is shown in green, and unfavorable closed-shell repulsions between the fluorine lone pair orbitals and the breaking σCO orbital (nF−σCO) are shown in red. The interplay of these effects will be discussed in the second part of this report for oxetenes 11−18. Scheme 2 shows the relative stability of trifluoromethylsubstituted oxetenes at ambient conditions. In contrast, other oxetenes undergo rapid electrocyclic ring opening to afford α,βunsaturated carbonyl compounds without isolation of the oxetene intermediates (Scheme 3).5a We have applied density functional theory to understand the origin of reactivities and torquoselectivities of mono-, di-, and trifluoromethyloxetenes. Scheme 4 shows the scope of cyclobutenes and oxetenes studied here. The electrocyclic ring opening reactions of oxetenes and cyclobutenes without alkenyl substituents (1−10), 2-phenyloxetenes bearing mono-, di-, and trifluoromethyl groups at the 2-position (11−18), and 2-phenyloxetenes disubstituted at the 2-positions were studied to determine the origin of the reactivities and torquoselectivities.



COMPUTATIONAL METHODS

All computations were carried out with Gaussian09, Rev.D01.12 Reactants, transition states, and products were optimized with the density functional M06-2X using the 6-31+G(d,p) basis set and the IEF-PCM13 model for toluene to correct for solvent effects. Vibrational analysis confirmed all stationary points to be minima (no imaginary frequencies) or transition structures (one imaginary frequency). Thermal corrections were computed from unscaled frequencies for the standard state of 1 atm and 298.15 K.

Scheme 1. 4π-Electrocyclic Ring Opening of Carbocyclic and Heterocyclic Four-Membered Rings



RESULTS AND DISCUSSION The lowest energy transition structures for both modes of conrotation for parent cyclobutenes (1−5) and oxetene derivatives (6−10) are given in Figure 1. For all transition structures, possible steric clashes are indicated by dotted black lines for contacts with radii less than the sum of their van der Waals radii.14 The disfavored transition structures are given in the Supporting Information with their activation free energies. The breaking C−C bond lengths range from 2.13 to 2.18 Å in the carbocyclic transition structures (TS1−TS5). The activation free energy (ΔG⧧) of cyclobutene is 35.2 kcal mol−1. The 3-methylcyclobutene is torquoselective for the “out” transition structure and has an activation free energy (ΔG⧧ = 32.6 kcal mol−1) lower than that of cyclobutene. The 3-trifluoromethylcyclobutene also prefers to rotate outward but is predicted to undergo electrocyclic ring opening much more slowly (ΔG⧧ = 36.6 kcal mol−1). The preference for the “out” transition state is due to the decreased steric congestion in that mode of conrotatory ring opening. The disubstituted cyclobutenes (3 and 5) have relatively bulky substituents (CF3 and CH3), and their transition states feature relatively long C−C and C−O breaking bond lengths (2.18 Å) to minimize closedshell repulsion between the substituents and the breaking σCC orbital. The unavoidable steric clashes in TS3 and TS5 result in relatively high activation free energies (37.4 and 38.2 kcal mol−1, respectively). The oxetene transition structures have C−O breaking bond lengths that range from 1.96 to 2.06 Å, somewhat shorter than those of the carbocyclic transition structures. The C−O breaking bond lengths are shorter in the reactant state, which are reflected in the transition structures. The oxetenes have

Thermal 4π-electrocyclic ring openings proceed via conrotation of the termini to form the corresponding dienes, in accordance with the Woodward−Hoffmann rules.11 3-Trifluoromethylcyclobutenes undergo slightly torquoselective electrocyclic ring opening reactions with inward rotation of the trifluoromethyl group.10 Mikami and co-workers report high torquoselectivities for the electrocyclic ring opening of 3trifluoroalkyloxetenes (>75% inward rotation product). These studies suggest that a through-space orbital interaction between the breaking σCO orbital with the σCF* orbital of the trifluoromethyl group is responsible for the preference of inward rotation in oxetenes. Scheme 2 illustrates this orbital interaction in transition structures where the fluoroalkyl group Scheme 2. Orbital Interactions upon Inward Rotation of the Fluoroalkyl Substituent

B

DOI: 10.1021/acs.joc.5b01361 J. Org. Chem. XXXX, XXX, XXX−XXX

Article

The Journal of Organic Chemistry

Downloaded by UNIV OF NEBRASKA-LINCOLN on September 2, 2015 | http://pubs.acs.org Publication Date (Web): August 28, 2015 | doi: 10.1021/acs.joc.5b01361

Scheme 3. Synthesis of Oxetenes via a [2 + 2] Cycloaddition and Subsequent Ring Opening Reaction9a

the parent and alkyl-substituted cases but lowest for the trifluoromethyl-substituted oxetenes. The electrocyclic ring openings of both 2-methyloxetene (7) and 2-trifluoromethyloxetene (9) prefer outward rotation of the substituent and have activation free energies of 25.1 and 32.5 kcal mol−1, respectively. 3,3-Disubstituted oxetenes (8 and 10) have relatively high activation barriers due to unavoidable closedshell substituent−σCO interactions in these transition states. In this section, we report transition structures for 2-phenylsubstituted oxetenes and evaluate how different fluoroalkyl groups affect the reactivities and torquoselectivities of oxetenes 11−15. The transition structures for both modes of conrotatory ring opening and corresponding activation free energies are shown in Figure 2. Reactivity. The 2-phenyloxetenes have activation free energies ∼1 kcal mol−1 higher than that of the parent oxetene. The breaking C−O bond length in these transition structures ranges from 1.95 to 2.02 Å; they are relatively unchanged from those of the TS6−TS10 without the phenyl. The 2monosubstituted oxetenes bearing substituents with increasing σ-acceptor characteristics have higher activation barriers along the series 12, 13, 14, and 15 (26.8−34.0 kcal mol−1). The 7.2 kcal mol−1 increase in activation free energy makes the ring opening of 15 approximately 500-fold slower than that for 12, which is in line with the elevated temperatures required for the ring opening of 15 (70 °C; Scheme 2). Torquoselectivity. Torquoselectivity is the difference in activation free energies between the two modes of conrotation (ΔΔG⧧). The electrocyclic ring opening of 2-methyloxetene 12 is predicted to be the most facile of the reactions shown in Figure 3. The ΔG⧧ of TS12-out is 26.8 kcal mol−1 and is predicted to be the most torquoselective for outward rotation of CH3 (ΔΔG⧧ = 5.8 kcal mol−1). The ΔG⧧ for the ring opening of monofluoromethyloxetene 13 is 2.2 kcal mol−1 higher and less torquoselective but still favors the outward rotation of CH 2 F by 2.6 kcal mol −1 . The decreased torquoselectivity stems from the stabilization of the “in” transition structure via a through-space interaction (σCO → σCF*) between the σCO orbital and the σCF* orbital (σCO → σCF*). The C−F bond in TS13-in is oriented to maximize σCO → σCF*. The reversal of torquoselectivity occurs when monofluoromethyl is changed to difluoromethyl; the TS14-in of the difluoromethyl case is increased by 3.2 kcal mol−1 compared to TS13-out. However, the TS14-in is now lower in energy than TS14-out by 1.4 kcal mol−1. Thus, the ring opening of 14 is torquoselective for inward rotation; the energy of the σCF*

Scheme 4. Mono- and Disubstituted Cyclobutene and Oxetene Derivatives

Figure 1. Optimized transition structures of model cyclobutenes and oxetenes. Distances and energies are reported in angstroms and kcal mol−1, respectively.

activation free energies that are 4−10 kcal mol−1 lower than those of the cyclobutene derivatives. The difference is large for C

DOI: 10.1021/acs.joc.5b01361 J. Org. Chem. XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF NEBRASKA-LINCOLN on September 2, 2015 | http://pubs.acs.org Publication Date (Web): August 28, 2015 | doi: 10.1021/acs.joc.5b01361

The Journal of Organic Chemistry

Figure 3. Transition structures for the electrocyclic ring opening reactions of disubstituted oxetenes 16−18. Distances and energies are reported in angstroms and kcal mol−1, respectively. Computed using M06-2X/6-31+G(d,p)/PCMtol.

coordinates, and energies of the other 2,2-disubstituted oxetenes can be found in the Supporting Information. The 2,2-disubstituted oxetenes have activation free energies generally higher than those of the 2-monosubstituted oxetenes. The ΔG⧧ values of TS16 and TS17 are 30.0 and 37.0 kcal mol−1, respectively. The ΔG⧧ values for 2-methyloxetene (12) and 2-trifluoromethyloxtene (15) are 26.8 and 34.0 kcal mol−1, respectively. These barriers are higher by ∼3 kcal mol−1 because unavoidable closed-shell repulsions (nF−σCO) make the transition states more unfavorable. The electrocyclic ring opening of 18 can involve either outward rotation of CH3 and inward rotation of CF3 (TS18-out) or inward rotation of CH3 and outward rotation of CF3 (TS18-out). The ΔG⧧ for the favored transition state (TS18-in) is 31.6 kcal mol−1. This ring opening is very torquoselective (ΔΔG⧧ = 6.1 kcal mol−1) because of simultaneously favorable outward rotation of CH3 and inward rotation of CF3. Inward rotation of CF3 facilitates σCO → σCF* while avoiding closed-shell repulsions between methyl hydrogens with σCO.

Figure 2. Two transition structures for conrotatory electrocyclic ring openings of 11−15. Distances and energies are reported in angstroms and kcal mol−1, respectively.

orbital is lower than that of 13 because of the additional geminal fluorine, and the σCO → σCF* becomes more favorable. The activation free energies of the trifluoromethyl case, TS15out and TS15-in, are both 34.0 kcal mol−1. This reaction is not torquoselective because TS15-in has a competition of the σCO → σCF* effect and unavoidable closed-shell repulsions between fluorine lone pair orbitals (nF) and (σCO) upon inward rotation. These interaction lengths are 2.63 and 2.76 Å, which are less than the sum of the oxygen and fluorine van der Waals radii (2.99 Å).14 The geometry of TS15-out does not allow for σCO → σCF* stabilization. An NBO analysis shows that the σCO → σCF* is present when the fluoroalkyl substituents rotate inward.9b We also explored a number of reactions of 2,2-disubstituted substrates. These substituents are mono-, di-, and trifluoromethyl groups like 12−15 (CH 2F, CHF2, and CF 3, respectively). The subsequent discussion will focus on the ring opening of 2,2-dimethyloxetene (16), 2,2-bistrifluoromethyloxetene (17), and 2-methyl-2-trifluoromethyloxetene (18). The transition structures and activation free energies for these reactions are given in Figure 3. The transition structures,



CONCLUSION The reactivities of oxetene electrocyclic ring openings are controlled by the electronic character of the substituent of 2substituted oxetenes. Oxetene ring opening is accelerated when the substituent at the 2-position is a σ-donor. σ-Acceptors cause the ring opening to occur less readily due to stronger nF−σCO interaction upon inward rotation of the fluoromethyl substituent, in particular. The torquoselectivities are thus controlled by a competition of σCO → σCF* effect and unfavorable closed-shell repulsions. Our calculations suggest that 2-difluoromethyloxetenes prefer inward rotation because the favorable σCO → σCF* orbital effect outweighs closed-shell repulsions in the transition state.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.5b01361. D

DOI: 10.1021/acs.joc.5b01361 J. Org. Chem. XXXX, XXX, XXX−XXX

Article

The Journal of Organic Chemistry



(10) (a) Lee, P. S.; Zhang, X.; Houk, K. N. J. Am. Chem. Soc. 2003, 125, 5072−5079. (b) Niwayama, S.; Kallel, A. K.; Spellmeyer, D. C.; Sheu, C.; Houk, K. N. J. Org. Chem. 1996, 61, 2813−2825. (11) Woodward, R. B.; Hoffmann, R. Angew. Chem., Int. Ed. Engl. 1969, 8, 781−932. (12) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision C.01; Gaussian Inc.: Wallingford, CT, 2010. (13) Cances, E.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 1997, 107, 3032. (14) Bondi, A. J. Phys. Chem. 1964, 68, 441.

Structures and energies of all considerable reactants and transition structures (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. Author Contributions §

K.H. and S.A.L. contributed equally to this paper.

Notes

The authors declare no competing financial interest.

Downloaded by UNIV OF NEBRASKA-LINCOLN on September 2, 2015 | http://pubs.acs.org Publication Date (Web): August 28, 2015 | doi: 10.1021/acs.joc.5b01361



ACKNOWLEDGMENTS This research was supported by Japan Science and Technology Agency (JST) (ACT-C: Creation of Advanced Catalytic Transformation for the Sustainable Manufacturing at Low Energy, Low Environmental Load). This work utilized the TSUBAME 2.5 supercomputer in Global Scientific Information and Computing Center (GSIC), Tokyo Institute of Technology. We thank the National Science Foundation for financial support of this work (CHE-1361104). Calculations were performed on the Hoffman2 cluster at UCLA and the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation Grant Number OCI-10535.



REFERENCES

(1) (a) Woodward, R. B.; Hoffmann, R. J. Am. Chem. Soc. 1965, 87, 395. (b) Vogel, E. Angew. Chem. 1954, 66, 640. (2) (a) Kirmse, W.; Rondan, N. G.; Houk, K. N. J. Am. Chem. Soc. 1984, 106, 7989. (b) Rondan, N. G.; Houk, K. N. J. Am. Chem. Soc. 1985, 107, 2099. (c) Jefford, C. W.; Bernardinelli, G.; Wang, Y.; Spellmeyer, D. C.; Buda, A.; Houk, K. N. J. Am. Chem. Soc. 1992, 114, 1157. (3) (a) Lopez, S. A.; Houk, K. N. J. Org. Chem. 2014, 79, 6189−6195. (b) Walker, M.; Hietbrink, B. N.; Thomas, B. E., IV; Nakamura, K.; Kallel, E. A.; Houk, K. N. J. Org. Chem. 2001, 66, 6669. (c) Mangelinckx, S.; Van Speybroeck, V.; Vansteenkiste, P.; Waroquier, M.; De Kimpe, N. J. Org. Chem. 2008, 73, 5481. (4) (a) Jaccob, M.; Jem, I. S.; Giri, S.; Venuvanalingam, P.; Chattaraj, P. K. J. Phys. Org. Chem. 2011, 24, 460. (b) Yu, H.; Chan, W.; Goddard, J. D. J. Am. Chem. Soc. 1990, 112, 7529. (5) (a) Kirsch, P. Modern Fluoroorganic Chemistry: Synthesis, Reactivity, Applications; Wiley-VCH: Weinheim, Germany, 2013. (b) Ojima, I. Fluorine in Medicinal Chemistry and Chemical Biology; Wiley-Blackwell: New York, 2009. (c) Begue, J. P.; Delpon, D. B. Bioorganic and Medicinal Chemistry of Fluorine; Wiley-Blackwell: New York, 2008. (6) (a) Wang, J.; Sánchez-Roselló, M.; Aceña, J. L.; del Pozo, C.; Sorochinsky, A. E.; Fustero, S.; Soloshonok, V. A.; Liu, H. Chem. Rev. 2014, 114, 2432−2506. (b) Purser, S.; Moore, P. R.; Swallow, S.; Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320−330. (7) O’Hagan, D. Chem. Soc. Rev. 2008, 37, 308−319. (8) The most popular ways to construct the stereoselective fluoromethyl-substituted alkene derivatives are stereospecific substitution to the sterically pure alkene precursors: (a) He, Z.; Luo, T.; Hu, M.; Cao, Y.; Hu, J. Angew. Chem., Int. Ed. 2012, 51, 3944−3947. (b) Zine, K.; Petrignet, J.; Thibonnet, J.; Abarbri, M. Synlett 2012, 23, 755−759. (c) Takeda, Y.; Shimizu, M.; Hiyama, T. Angew. Chem., Int. Ed. 2007, 46, 8659−8661. (d) Liu, X.; Shimizu, M.; Hiyama, T. Angew. Chem., Int. Ed. 2004, 43, 879−882. (e) Aikawa, K.; Nakamura, Y.; Yokota, Y.; Toya, W.; Mikami, K. Chem. - Eur. J. 2015, 21, 96−100. (9) (a) Aikawa, K.; Hioki, Y.; Shimizu, N.; Mikami, K. J. Am. Chem. Soc. 2011, 133, 20092−20095. (b) Aikawa, K.; Shimizu, N.; Honda, K.; Hioki, Y.; Mikami, K. Chem. Sci. 2014, 5, 410−415. E

DOI: 10.1021/acs.joc.5b01361 J. Org. Chem. XXXX, XXX, XXX−XXX